Controlled Polymerization of a Cyclic Diene Prepared from the Ring

May 20, 2009 - The synthesis of polymers from renewable resources is driven by the desire to reduce our dependence on petroleum-based products...
1 downloads 0 Views 150KB Size
Published on Web 05/20/2009

Controlled Polymerization of a Cyclic Diene Prepared from the Ring-Closing Metathesis of a Naturally Occurring Monoterpene Shingo Kobayashi, Cheng Lu, Thomas R. Hoye,* and Marc A. Hillmyer* Department of Chemistry, UniVersity of Minnesota, Minneapolis, Minnesota 55455-0431 Received April 6, 2009; E-mail: [email protected]; [email protected]

The synthesis of polymers from renewable resources is driven by the desire to reduce our dependence on petroleum-based products and to develop sustainable materials technologies.1 Much effort has focused on carbohydrate-derived hydroxy acid or vegetable oil based starting materials. Naturally occurring terpenes include unsaturated hydrocarbons and comprise another attractive class of biobased polymer precursors.2 The importance of terpenes in the sustainable materials arena is underscored by the recent announcement of the planned commercial scale production of isoprene by fermentation.3 The monoterpene myrcene (1) can be readily obtained from plants4 or from the pyrolysis of pinene.5 We expected that the cyclic diene 3-methylenecyclopentene (2) could be prepared from 1 by ring-closing metathesis (RCM)6 and that 2 would be an attractive substrate for polymerization (Scheme 1). Furthermore, Scheme 1

the byproduct of the metathesis cyclization of 2 is isobutene, the principal starting material for butyl rubber. Here, we report the synthesis of 2 using RCM and of poly-2 using radical, anionic, and cationic polymerization. We demonstrate that the controlled cationic polymerization of 2 gives regiopure 1,4-poly-2 and the hydrogenation of this polymer yields regiopure poly(cyclopentane1,3-diylmethylene).7 We also show that 1,4-poly-2 is semicrystalline by differential scanning calorimetry (DSC) with a melting temperature as high as 105 °C. Compound 1 was converted to 2 in 45% yield (68% conversion) by RCM using 0.2 mol % of the Grubbs second generation initiator [G2, H2IMes(Cl2)(Cy3P)RudCHPh] in decalin at 40 °C for 5 h. Complete conversion of 1 to 2 was achieved using 1.0 mol % G2. Due to its low concentration in the reaction mixture and lack of ring-strain, ring-opening metathesis polymerization of 2 is unfavorable. The RCM synthesis of 2 can be carried out on gram scale and represents a rare example of an ene-diene RCM reaction leading to an exocyclic conjugated double bond.8 The secondary product isobutene is also polymerizable and enhances the atom economy of our overall strategy. Previous reports describing the synthesis of 2 are not of preparative value.9 Radical,10 anionic,11 and cationic12 polymerizations of 2-methyl1,3-pentadiene, an acyclic structural analogue of 2, are known. In each case, the polymer obtained was of mixed regio- and stereochemistry. We studied the polymerization of 2 using the radical initiator AIBN. The conversion in benzene solution was low, but bulk polymerization afforded a polymeric product in 58% yield after 20 h at 80 °C. The Mn values of the resulting polymers were less than 1 kg mol-1 by SEC versus PS standards. Anionic polymerization of 25 equiv of 2 in cyclohexane using secbutyllithium (s-BuLi) as the initiator, on the other hand, afforded 7960

9

J. AM. CHEM. SOC. 2009, 131, 7960–7961

polymeric products in near quantitative conversion after 96 h at 70 °C or after 5 h at 40 °C in the presence of tetramethylethylenediamine ([TMEDA]0/[s-BuLi]0 ) 2).13 In both cases the poly-2 product exhibited a narrow molecular weight distribution by SEC vs PS standards (PDI < 1.2). Without TMEDA the degree of polymerization calculated by 1H NMR spectroscopy assuming one s-butyl initiating group per chain was 24, consistent with the predicted value of 25 (Table S1). Cationic initiating systems were also effective for inducing polymerization of 2. Upon addition of BF3 · OEt2 (RT, 5 min) or Me3O+BF4- (40 °C, 20 min) the polymerization media became highly viscous, and the conversion of 2 was quantitative in both cases (Table S1). The BF3 · OEt2-initiated sample of poly-2 exhibited quite low solubility, but the polymer obtained with Me3O+BF4possessed good solubility in common organic solvents. Both samples of poly-2 exhibited low molecular weights (Mn < 4 kg mol-1) and broad molecular weight distributions by SEC. We also explored living cationic polymerizations with the known i-BuOCH(Cl)Me/Lewis acid initiator systems (Table 1).14 These Table 1. Cationic Polymerization of 2 with i-BuOCH(Cl)Me/Lewis

Acid/Et2O in Toluene

Mn (kg mol-1)

entry

Lewis acid

time (min)

conva (%)

calcdb

NMRa

PDIc

Tg (°C)

Tm (°C)

1d 2e 3e 4f 5g

SnCl4 ZnCl2 ZnCl2 ZnCl2 ZnCl2

1 1 2 5 9

100 68 85 88 94

5.5 3.8 4.7 9.5 21

7.7 3.8 4.7 8.7 22

6.51 1.25 1.15 1.15 1.21

-4 -8 -8 0 11

68, 101 71-87h 66, 87 66, 101 65, 105

a1 H NMR spectroscopy using the methyl group from the initiator (see Supporting Information). b Mncalcd ) (MW of monomer) × M/I × conversion + MW of initiator residue. c SEC in CHCl3. d 2/ i-BuOCH(Cl)Me/SnCl4/Et2O ) 68:1:1:20 in toluene (5.2 mL) at -78 °C. e 2/i-BuOCH(Cl)Me/ZnCl2 ) 68:1:1, in Et2O/toluene solution (0.8 mL/4.4 mL) -40 °C. f 2/i-BuOCH(Cl)Me/ZnCl2 ) 136:1:1, in Et2O/ toluene solution (0.44 mL/6.3 mL) -40 °C. g 2/i-BuOCH(Cl)Me/ZnCl2 ) 272:1:1, in Et2O/toluene solution (0.3 mL/8.2 mL) -40 °C. h Multiple melting transitions were observed for this sample.

polymerizations were rapid at low temperature. While the poly-2 obtained using i-BuOCH(Cl)Me/SnCl4/Et2O15 had a broad molecular weight distribution, the i-BuOCH(Cl)Me/ZnCl2/Et2O16 system afforded polymers with narrow molecular weight distributions. The Mn values for poly-2 increased with the conversion of 2 and were in good agreement with the calculated values (Table 1, entries 2-5). Furthermore, the apparent Mn values could be tuned by changing the feed molar ratios of monomer to initiator. Thus, the cationic polymerization of 2 proceeded in a controlled fashion. Figure 1 shows typical 1H NMR spectral features of the vinylic protons in poly-2 samples obtained by anionic and cationic 10.1021/ja9027567 CCC: $40.75  2009 American Chemical Society

COMMUNICATIONS

1 H NMR spectra of poly-2 prepared with (a) s-BuLi/TMEDA, (b) s-BuLi, and (c) i-BuOCH(Cl)Me/ZnCl2/Et2O.

Figure 1.

polymerizations. For the former, two sets of signals at ∼5.3 and 4.8 ppm were assigned to repeating units with 1,4- (Ha) and 4,3regiochemistry (Hb/Hc), respectively. As with other anionic polymerizations of dienes, the use of TMEDA leads to a high content of 4,3-repeating units [96%, panel (a)]. Higher 1,4-content was observed without TMEDA [74%, panel (b)]. The 1H NMR [panel (c)] and 13C NMR (Figure 2) spectra of poly-2 samples prepared

Figure 2. 13C NMR spectrum of 1,4-poly-2 prepared with i-BuOCH(Cl)Me/ ZnCl2/Et2O (Assignments are based on analyses of COSY, HMQC, and DEPT NMR experiments).

by cationic polymerization were consistent with essentially regiopure structures. Cationic polymerization of the analogous 3-methylene cyclohexene also gives regiopure material.12 In samples of poly-2 prepared using the i-BuOCH(Cl)Me/Lewis acid system, low intensity resonances in the 1H NMR spectra (5.8-6.5 ppm) suggested the presence of a cyclopentadienyl group (cf. CPD in Table 1 graphic) at the chain end. Reaction of an oligomeric sample with maleic anhydride afforded Diels-Alder adducts (Figure S10a). We also showed that poly-2-b-polylactide block copolymers can be formed by the Diels-Alder coupling of a maleimide end-functionalized polylactide (Figure S10b). Thus the reactive CDP end group in poly-2 can be exploited for a variety of end-functionalization reactions. We could hydrogenate 1,4-poly-2 and convert it into the saturated variant, poly(cyclopentane-1,3-diylmethylene), using p-toluenesulfonylhydrazide. SEC data for the hydrogenated product showed a narrow and unimodal peak at slightly lower elution volume consistent with the absence of significant side reactions (Figure S17). The 13C NMR spectrum (Figure S14) of hydrogenated poly-2 is consistent both with the spectrum of stereoirregular poly(cyclopentane-1,3-diylmethylene), synthesized by the cyclopolymerization of 1,5-hexadiene,7 and with our conclusion that poly-2 formed by cationic polymerization is regiopure. We analyzed the thermal properties of poly-2 and hydrogenated poly-2 by DSC. High 1,4-poly-2 exhibited Tg values between -17 and 11 °C depending on molecular weight (Table 1). High 4,3poly-2 exhibited a particularly high Tg of 73 °C at relatively low molecular weight (Mn ) 3.6 kg mol-1). 1,4-Poly-2 from the i-BuOCH(Cl)Me/Lewis acid/Et2O systems exhibited multiple endo-

thermic peaks that we assign to melting and/or liquid crystalline transitions (Table 1).17 Regioregular poly-2 from the BF3 · OEt2 and Me3O+BF4- systems did not show endothermic peaks (Figure S18). Such semicrystallinity suggests that regioregular poly-2 obtained using i-BuOCH(Cl)Me/Lewis acid/Et2O may have a more stereoregular structure. Hydrogenated 1,4-poly-2 exhibited a Tg of -28 °C and a Tm of 106 °C, consistent with previously reported values for trans-rich or random poly(cyclopentane-1,3-diylmethylene).17 The regiopurity and narrow molecular weight distribution of hydrogenated 1,4-poly-2 will facilitate systematic studies on the recently reported liquid crystalline behavior of this polyolefin.17 In conclusion, we demonstrated the RCM conversion of the natural monoterpene myrcene (1) to 3-methylenecyclopentene (2). Monomer 2 can be polymerized under radical, anionic, or cationic conditions. Cationic polymerization using the i-BuOCH(Cl)Me/ ZnCl2/Et2O system afforded regiopure 1,4-poly-2 with controlled molecular weight and narrow molecular weight distribution. Exploitable 1,3-diene end groups were identified in 1,4-poly-2. Finally, 1,4-poly-2 exhibited melting transitions consistent with a stereoregular structure. Acknowledgment. Primary support from the NSF MRSEC Program (DMR-0212302, DMR-0819885). We thank Prof. W. B. Tolman for helpful input, W. Gramlich for maleimide functionalized polylactide, Prof. S. R. Kass, M. Meyer, and J. Schmidt for MS assistance and Materia, Inc. for donation of G2. Supporting Information Available: Experimental details and characterization data. This material is available free of charge via the Internet at http://pubs.acs.org. References (1) (a) Mecking, S. Angew. Chem., Int. Ed. 2004, 43, 1078–1085. (b) Williams, C. K.; Hillmyer, M. A. Polym. ReV. 2008, 48, 1–10. (c) Gandini, A. Macromolecules 2009, 41, 9491–9504. (2) (a) Wanamaker, C. L.; O’Leary, L. E.; Lynd, N. A.; Hillmyer, M. A.; Tolman, W. B. Biomacromolecules 2007, 8, 3634–3640. (b) Lu, J.; Kamigaito, M.; Sawamoto, M. Macromolecules 1997, 30, 22–26. (3) McCoy, M. Chem. Eng. News 2008, 86 (38), 15. (4) Zoghbi, M. G. B.; Andrade, E. H. A.; Da Silva, M. H. L.; Carreira, L. M. M.; Maia, J. G. S. FlaVour Fragr. J. 2003, 18, 421–424. (5) (a) Burwell, R. L., Jr. J. Am. Chem. Soc. 1951, 73, 4461–4462. (b) Erman, W. E. Chemistry of the Monoterpenes: An Encyclopedic Handbook; Marcel Dekker: New York, 1985. (6) Grubbs, R. H.; Miller, S. J.; Fu, G. C. Acc. Chem. Res. 1995, 28, 446–452. (7) (a) Coates, G. W.; Waymouth, R. M. J. Am. Chem. Soc. 1993, 115, 91–98. (b) de Ballesteros, O. R.; Venditto, V.; Auriemma, F.; Guerra, G.; Resconi, L.; Waymouth, R.; Mogstad, A. L. Macromolecules 1995, 28, 2383–2388. (8) For example see: (a) Paquette, L. A.; Tae, J.; Arrington, M. P.; Sadoun, A. H. J. Am. Chem. Soc. 2000, 122, 2742–2748. (b) Kirkland, T. A.; Grubbs, R. H. J. Org. Chem. 1997, 62, 7310–7318. (9) For example see: Adam, W.; Alt, C.; Braun, M.; Denninger, U.; Zang, G. J. Am. Chem. Soc. 1991, 113, 4563–4571, and additional references in the Supporting Information. (10) Kamachi, M.; Kajiwara, A. Macromolecules 1996, 29, 2378–2382. (11) Xu, Z.; Mays, J.; Chen, X.; Hadjichristidis, N.; Schilling, F. C.; Bair, H. E.; Pearson, D. S.; Fetters, L. J. Macromolecules 1985, 18, 2560–2566. (12) Imanishi, Y.; Kanagawa, S.; Higashimura, T. Makromol. Chem. 1974, 175, 1761–1776. (13) (a) Natori, I. Macromolecules 1997, 30, 3696–3697. (b) Hong, K.; Mays, J. W. Macromolecules 2001, 34, 782–786. (14) (a) Miyamoto, M.; Sawamoto, M.; Higashimura, T. Macromolecules 1984, 17, 265–268. (b) Schappacher, M.; Deffieux, A. Macromolecules 1991, 24, 2140–2142. (15) (a) Ouchi, M.; Kamigaito, M.; Sawamoto, M. Macromolecules 2001, 34, 3176–3181. (b) Mizuno, N.; Satoh, K.; Kamigaito, M.; Okamoto, T. Macromolecules 2006, 39, 5280–5285. (16) Sawamoto, M.; Okamoto, C.; Higashimura, T. Macromolecules 1987, 20, 2693–2697. (17) Naga, N.; Yabe, T.; Sawaguchi, A.; Sone, M.; Noguchi, K.; Murase, S. Macromolecules 2008, 41, 7448–7452.

JA9027567

J. AM. CHEM. SOC.

9

VOL. 131, NO. 23, 2009

7961